검색 전체 메뉴
PDF
맨 위로
OA 학술지
Phylogeographic and Feeding Ecological Effects on the Mustelid Faunal Assemblages in Japan
  • 비영리 CC BY-NC
  • 비영리 CC BY-NC
ABSTRACT
Phylogeographic and Feeding Ecological Effects on the Mustelid Faunal Assemblages in Japan
KEYWORD
phylogeny , phylogeography , feeding ecology , species assemblage , Mustelidae
  • INTRODUCTION

    Endemism and vulnerability are two important sides of island biodiversity. Island organisms show peculiar morphologies (Millien, 2006) and genetic properties (Sato et al., 2009b), while small populations on islands with such endemism tend to be fragile due to lower genetic potential and/or environmental stochasticity (Frankham et al., 2010). Kier et al. (2009) assessed endemism richness of plants and vertebrates worldwide and clarified that the island biota is more endemic and susceptible to the human impact than the continental ones. Species facing imminent extinction was shown to be concentrated on island biota (Ricketts et al., 2005). Islands can therefore appropriately be called “biodiversity and extinction hotspots.” Understanding the mechanisms for the gain and loss of the biodiversity of the islands is imperative for the conservation of the endangered species that is being lost on islands.

    The Japanese archipelagos consist of more than 6,800 islands. They are usually arranged into three biogeographic regions shown on Fig. 1: Hokkaido, Honshu-Shikoku-Kyushu, and Ryukyu (Kawamura, 1991, 1994; Dobson, 1994; Dobson and Kawamura, 1998). The major islands, which are supposed to have been repeatedly connected to the Asian continent during the Pleistocene glacial periods, have received influxes of continental organisms and isolated them differentially in each biogeographic region (Dobson, 1994; Millien-Parra and Jaeger, 1999), leading to various levels of endemic organisms among major islands (Suzuki, 2009). Compared with the same

    latitudinal regions, the Japanese archipelagos have rich endemic species, especially in mammalian fauna (Kier et al., 2009). The Japanese islands provide an excellent opportunity for understanding the mechanisms for the generation of the biodiversity.

    In the modern conservation framework, evolutionary distinctiveness is considered useful for the conservation prioritization (Isaac et al., 2007). Evolutionary history should be clarified to know what should be conserved first (Collen et al., 2011). In addition, adaptive ecological characteristics should also be considered in conservation biology (Ryder, 1986; de Guia and Saitoh, 2007; Isaac et al., 2007; Allendorf et al., 2010; Collen et al., 2011; Funk et al., 2012). It is therefore of great importance to clarify the contribution of both phylogenetic and ecological effects to the formation of species assemblages underpinning regional biodiversity (Webb et al., 2002; Wiens and Donoghue, 2004; Wiens and Graham, 2005).

    This review summarizes evolutionary histories and food habits of mustelid carnivoran species in Japan, and assesses the historical and ecological causes for the community assemblage of mustelid fauna in the different biogeographic regions.

    There are seven extant and indigenous terrestrial mustelid species in Japan (Ohdachi et al., 2009), the Japanese marten Martes melampus (Wagner, 1840), the sable Martes zibellina (Linnaeus, 1758), the Japanese badger Meles anakuma Temminck, 1844, the ermine or the stoat Mustela erminea Linnaeus, 1758, the Japanese weasel Mustela itatsi Temminck, 1844, the least weasel Mustela nivalis Linnaeus, 1766, and the Siberian weasel Mustela sibirica Pallas, 1773. The natural distributions of these seven species (Fig. 2) can be summarized into four biogeographic patterns: 1) present only in Hokkaido in Japan with the same species expanded to the continental Palaearctic region (Ma. zibellina) (Fig. 2A); 2) endemic to the Honshu-Shikoku-Kyushu region (Ma. melampus, Mu. itatsi, and Me. anakuma) (Fig. 2B-D); 3) present only in Tsushima Islands (Mu. sibirica) (Fig. 2D); and 4) Present in Hokkaido and a part of the northern Honshu in

    Japan with the same species expanded to the Holarctic region (Mu. erminea and Mu. nivalis) (Fig. 2E, F). Recent extensive accumulation of data on evolutionary and feeding ecological studies would enable clarification of the historical and ecological meaning of the distributions.

    I first reviewed the molecular evolutionary and feeding ecological studies on mustelids in Japan, and then discuss the faunal assemblage of the Japanese mustelid species. Two species below were not included in this review. The American mink Neovison vison (Schreber, 1777) is considered an exotic species from North America (Uraguchi, 2009). The Eurasian otter Lutra lutra (Linnaeus, 1758), an aquatic mustelid, is currently recognized as an extinct species (Sasaki, 2009b) and has provided too few phylogeographic and ecological data to review. In addition, I do not discuss the artificially introduced populations (e.g., Ma. melampus in Hokkaido and Sado Islands, Mu. itatsi in Hokkaido Islands, and Mu. sibirica in the western part of the Honshu Islands) to assess the establishment process of the natural distribution.

    BIOGEOGRAPHIC HISTORY OF THE MUSTELIDS IN JAPAN

      >  Biogeographic pattern I

    Martes zibellina. Recent molecular phylogenetic studies demonstrated that Ma. zibellina is closely related to the pine marten Martes martes (Linnaeus, 1758) (e.g., Koepfli et al., 2008; Sato et al., 2012) (Fig. 3). The divergence between them was estimated to have occurred approximately 1.0 million years ago (Mya) by chronological analyses using 22 nuclear and mitochondrial genes (Koepfli et al., 2008) (Fig. 3). Among intraspecific variations, the sable in Hokkaido is classified as a local subspecies, Martes zibellina brachyura (Temmink, 1844), based on morphology (Anderson, 1970; Wozencraft, 2005). Consistent with the morphological data,

    Sato et al. (2011) examining the mitochondrial NADH dehydrogenese subunit 2 (Nd2) gene sequences indicated that the Hokkaido sable possessed endemic genetic property. The analyses of the mitochondrial cytochrome b (Cytb) gene also supported the distinctive nature of the Japanese sable (Hosoda et al., 1999, 2000; Ishida et al., 2013). Sato et al. (2011) estimated the migration time of Ma. zibellina in Hokkaido to be 0.27-0.10 Mya, but stressed that 0.10 Mya would be closer to the migration date by considering the geological evidence. McKay (2012) re-examined the Cytb data of Hosoda et al. (2000) and provided 0.08 Mya as divergence time for Hokkaido and the continental populations. More recently, Ishida et al. (2013) examined two closely linked nuclear genes, melanocortin 1 receptor (Mc1r) and transcription factor 25 (Tcf25), and inferred the migration time to be ca. 0.05-0.01 Mya by using the recombination rate between these two genes. Although the time estimates above differed from each other, they consistently showed that the migration of this species from the continent could have occurred in the Late Pleistocene (0.13-0.01 Mya).

      >  Biogeographic pattern II

    Martes melampus. Although Kuroda and Mori (1923) reported the existence of this species in South Korea and some studies have provided corroboration (e.g., Wozencraft, 2005; Monakhov, 2011), I presently suppose that this species is endemic to Japan according to a recent illustrated reference book (Ohdachi et al., 2009). Previous interspecific systematic studies have clarified the species level endemism of this species, but failed to resolve the phylogenetic relationships among Ma. melampus, the American marten Martes americana (Turton, 1806), and a clade including Ma. martes and Ma. zibellina (Koepfli et al., 2008; Sato et al., 2012) (Fig. 3). Nevertheless, they consistently showed that the lineage of Ma. melampus was earlier generated than Ma. zibellina. The divergence time of Ma. melampus from the other Martes species was estimated to be 1.6-1.8 Mya (Koepfli et al., 2008; Sato et al., 2012) (Fig. 3). McKay (2012) re-examined the Cytb data of Hosoda et al. (2000) and showed a relatively younger estimate (1.36 Mya). It has not been clearly understood whether the ancestral lineage of Ma. melampus migrated from northern Sakhalin or southern Korean Peninsula. However, considering the current circumboreal distribution of the Holarctic martens in Eurasia, the migration through northern route via Sakhalin seems plausible. Recently Ishida et al. (2013) found some Mc1r and Tcf25 gene haplotypes from Ma. zibellina in Hokkaido that were closely related to those of Ma. americana and Ma. melampus. They might be considered remains of ancestral populations in Hokkaido leading to the extant Japanese marten lineages in Honshu Islands (Ishida et al., 2013).

    For intraspecific diversity of Ma. melampus, Kurose et al. (1999) and Inoue et al. (2010) did not detect any correlations between sampling locations and genetic variations in the mitochondrial Cytb and Dloop sequences, respectively. On the other hand, Sato et al. (2009b), by using more diverse data from three mitochondrial genetic loci (Cytb, Nd2, and Dloop), reported some extent of geographic structures within this species, showing distinct genetic property of each Tohoku, Tochigi-Niigata, Kyushu, Tsushima, and the other western Japanese lineage. To date, no studies have introduced time axis for intraspecific diversity of this species. Here, I simply calculated the within-species divergence times on the basis of the molecular clock assumption and 1.6-1.8 million years for the Ma. melampus-Ma. zibellina split, using data of Sato et al. (2009b). The result showed that the estimated time to the most recent common ancestor of the total examined individuals for Ma. melampus in Sato et al. (2009b) was 0.32- 0.36 Mya, indicating that the divergence events within this species might have occurred in the Middle Pleistocene (0.78- 0.13 Mya).

    Meles anakuma. The Japanese badger was once a subspecies of the Eurasian badger Meles meles (Linnaeus, 1758) (e.g., Wozencraft, 1993). However recent molecular phylogenetic studies have demonstrated that it has a distinctive evolutionary lineage from European and the other Asian continental badgers (Marmi et al., 2006; Cerro et al., 2010; Tashima et al., 2011). It is now regarded as a distinct species within the genus Meles together with European Me. meles and North and East Asian Meles leucurus (Hodgson, 1847) (see Wozencraft, 2005). Recently Cerro et al. (2010) introduced an additional Meles species, Me. canescens, from Southwest Asia on the basis of the genetic variations in six nuclear and one mitochondrial gene sequences. The molecular systematic conclusion is consistent with considerable geographic variations in morphology such as facial colour patterns and penis bone structures across the Eurasian continent and Japan (Abramov, 2002, 2003). Meles anakuma is the most closely related to Me. leucurus, and the divergence between them was estimated to have occurred 0.5 Mya (Tashima et al., 2011) (Fig. 3). The time estimate is in agreement with the fossil records, suggesting that the immigration through Korean Peninsula occurred in the middle Middle Pleistocene period (0.43 Mya) (Ogino et al., 2009). Kaneko (2009) supposed on the basis of the absence of this species in Hokkaido and Ryukyu Islands that an ancestral lineage of Me. anakuma migrated via Korean Peninsula. It should be noted that McKay (2012) re-examined the Cytb data of Kurose et al. (2001) and provided an older estimate (1.4 Mya) for the divergence time between Me. anakuma and Me. leucurus, which is inconsistent with the aforementioned molecular and fossil evidence (Ogino et al., 2009; Tashima et al., 2011).

    Kurose et al. (2001) showed very low intraspecific genetic diversity in the Cytb gene, and attributed the result to the recent population expansion. Tashima et al. (2011) supported Kurose et al. (2001) for the recent expansion by showing negative Tajima’s D (Tajima, 1989) and Fu’s Fs (Fu, 1997) values using the Dloop sequences. The population expansion was suggested to have begun 0.23-0.11 Mya (Tashima et al., 2011).

    Mustela itatsi. Although previously the Japanese weasel was classified as a subspecies of the Siberian weasel Mustela sibirica (e.g., Wozencraft, 1993), interspecific level of distinctiveness of this species has been supported by the recent molecular phylogenetic (e.g., Masuda and Yoshida, 1994; Sato et al., 2003, 2012) and morphological (Suzuki et al., 2011) studies. The Japanese weasel currently has a valid species name, Mustela itatsi (Wozencraft, 2005). Phylogenetic position of this species has remained to be elucidated. Sato et al. (2012) examined 10 genetic loci (nine nuclear and one mitochondrial genes), but failed to resolve the phylogenetic relationships among Mu. itatsi, Mu. sibirica, and the clade encompassing the European mustelids, the steppe polecat Mustela eversmanii Lesson, 1827, the European mink Mustela lutreola (Linnaeus, 1761), and the European polecat Mustela putorius Linnaeus, 1758 (Fig. 3). The molecular chronological analyses in Sato et al. (2009a, 2012) showed that the time for the trichotomous divergence among those three lineages was 1.5-1.6 Mya (Fig. 3). Although it is not clear when the lineage of Mu. itatsi placed its origin in Japan, Ogino et al. (2009) showed on the basis of the fossil evidence that Mu. itatsi was present in Kyushu Islands in the middle Middle Pleistocene period (0.43 Mya) at the latest. McKay (2012) re-examined the Cytb data of Hosoda et al. (2000) and Marmi et al. (2004), and showed more ancient divergence between Mu. itatsi and Mu. sibirica (2.88 Mya), which seems an overestimate compared with the molecular estimate from nine nuclear and one mitochondrial gene sequences (Sato et al., 2009a, 2012) and fossil evidence (Ogino et al., 2009).

    Few studies have contributed to the intraspecific genetic diversity of this species, but recently Masuda et al. (2012) summarized its phylogeographic history. By using the Dloop sequence variations, they clarified two distinct evolutionary lineages of the Japanese weasel, Honshu and Kyushu-Shikoku lineages. The divergence time of these two lineages was estimated to be 0.83-1.17 Mya. Their demographic analyses captured the tendency of the recent expansion of the Honshu population that occurred 0.54 or 0.77 Mya. On the other hand, the Kyushu-Shikoku population appeared to have more stable population history within refugia in the Pleistocene glacial periods, and the population expansion (albeit not clearly indicated) was inferred to have occurred 0.59 or 0.83 Mya.

      >  Biogeographic pattern III

    Mustela sibirica. The Siberian weasel is only observed in Tsushima Islands in Japan, providing a peculiar biogeographic pattern in the Japanese mustelid fauna, although descendants of the populations artificially introduced in the early 20th century are currently expanded to the western part of Japan (Sasaki, 2009a). As described above, the lineage of the Siberian weasel, Mu. sibirica, originally occurred in the early Pleistocene at the same time with that of the Japanese weasel, Mu. itatsi (1.5-1.6 Mya) (Fig. 3) (Sato et al., 2012).

    On the other hand, the migration time to the Tsushima Islands seems to be very recent because Mu. sibirica in Tsushima Islands was demonstrated to have the same Cytb gene haplotype (402 bp) as that observed in Korea (Hosoda et al., 2000). Meanwhile, Masuda et al. (2012) showed the genetic differentiations of the Dloop haplotypes observed in the Tsushima and Taiwan Islands from the continental Russian and Korean haplotypes. Since they did not infer the divergence times between these continental and island haplotypes, the chronology for the origin of the Tsushima populations has remained to be enlightened.

      >  Biogeographic pattern IV

    Mustela erminea. Phylogenetic analyses have indicated an origin of the lineage of Mustela erminea in the Early to Late Pliocene (3.5-4.0 Mya) (Fig. 3) (Koepfli et al., 2008; Sato et al., 2009a, 2012). In contrast to the relatively old origin compared with the other terrestrial mustelid species living in Japan, mitochondrial Dloop sequence analyses showed low intraspecific variations, suggesting recent diversifications across the Eurasian continent (Kurose et al., 2005). Meanwhile, Kurose et al. (2005) clarified that Mu. erminea in Japan had two distinct lineages, in which the clade including three individuals from both Honshu and Hokkaido showed close affinity to the North American individuals, while one Honshu individual was more closely related to a continental Russian individual than the other Japanese ones, implying multiple migrations into the Japanese archipelagos. Although time scale for intraspecific divergence has not been precisely estimated, fossil records in Kawamura et al. (1989) showed that Mu. erminea was present in the middle and late Middle Pleistocene in Yamaguchi prefecture (southern Honshu), and in the early Late Pleistocene in the Tochigi prefecture (central Honshu).

    Mustela nivalis. Mustela nivalis showed close phylogenetic affinity to the mountain weasel Mustela altaica Pallas, 1811, and divergence time between the lineages of these species was estimated to be 2.0-2.6 Mya in Late Pliocene to Early Pleistocene (Fig. 3) (Koepfli et al., 2008; Sato et al., 2009a, 2012). Although Mu. nivalis has a similar distribution pattern to Mu. erminea, intraspecific diversity is much more extensive for Mu. nivalis in terms of both morphological (Abramov and Baryshnikov, 2000) and genetic (Kurose et al., 2005) characteristics. As in Mu. erminea, Kurose et al. (2005) found two distinct lineages in Japan (Honshu and Hokkaido lineages), similarly requiring multiple migrations for their distributions. Since the karyotypes of the Honshu and Hokkaido populations differ from each other (2n=38 and 42, respectively), Obara (1991) proposed that the Honshu population be treated as distinct species, Mustela namiyei Kuroda, 1921. However, the level of the difference between Honshu and Hokkaido populations is currently considered an intraspecific variation (Hosoda et al., 2000; Kurose et al., 2005). To date, there have been no chronological analyses for the migrations into the Japanese archipelagos. Thus, the fossil discovered in the late Late Pleistocene in Gifu prefecture (central Honshu) is the only valuable chronological evidence (Kawamura et al., 1989).

    FEEDING ECOLOGY OF MUSTELIDS IN JAPAN

      >  Martes zibellina

    The sable is an arboreal carnivoran species in the genus Martes, occupying both coniferous and deciduous forests across northern Eurasia and Hokkaido, Japan (Murakami,

    [Table 1.] Dietary habits of seven extant mustelid species in Japan

    label

    Dietary habits of seven extant mustelid species in Japan

    2009). The sable largely depends on these forest resources. Martes species is usually considered an opportunistic generalist, whose diet changes depending on the environments they inhabit (Zhou et al., 2011). Meanwhile, the prey of Ma. zibellina is relatively biased to small mammals such as rodents (Zhou et al., 2011). Murakami (2003) examined food habits of the Japanese sable, and clarified that the sable preyed mainly on mammals like voles throughout the year, although it showed omnivorous nature depending on the availability of food items in different seasons. The sable can be considered an opportunistic specialist of rodents (Table 1).

      >  Martes melampus

    There is a tendency of Martes species to prey on mammals such as rodents in the colder northern region, while they eat more invertebrates and plants in warmer southern region (Zhou et al., 2011). Concordantly, Ma. melampus, an inhabitant of the southern Japanese islands, consumes more insects and fruits in its diet in addition to small mammals than most other martens (Shiratsuki, 1972; Suzuki et al., 1976, 1977; Yamagishi, 1990; Tatara and Doi, 1994; Nakamura et al., 2001; Arai et al., 2003; Ueuma et al., 2005; Zhou et al., 2011; Koike et al., 2012). This species is well known as a seed disperser of many plant species (Otani, 2002; Tsuji et al., 2011). Around suburban areas of Tokyo, seeds, fruits, and insects dominate as the main items in its fecal composition (Nakamura et al., 2001). In the Tsushima Islands, Ma. melampus preys on insects, centipedes, and plants more frequently than sympatric carnivores, the Siberian weasel Mu. sibirica and the leopard cat Prionailurus bengalensis (Kerr, 1792) (see Tatara and Doi, 1994). Ma. melampus tends to show very high diversity of food items, representing an opportunistic generalist property (Tatara and Doi, 1994; Yamamoto, 1994) (Table 1). Arai et al. (2003) mentioned that a wide variety of food items and flexible feeding strategy of Ma. melampus suggested high potential of adaptability to various food environments.

      >  Meles anakuma

    Badgers in the genus Meles are a morphologically and ecologically distinct form within the family Mustelidae, which have adapted to a fossorial life-style (Fig. 3) (Kaneko, 2009). The diet of Me. anakuma concentrates on earthworms from spring to autumn, although it also prefers fruits such as the persimmon in autumn (Yamamoto, 1991; Kaneko et al., 2006; Kaneko, 2009). Me. anakuma in Mt. Nyugasa (Nagano Prefecture) has the least diverse diet, being concentrated on earthworms and insects, compared with the other sympatric carnivores, the Japanese marten Ma. melampus, the red fox Vulpes vulpes (Linnaeus, 1758), and the raccoon dog Nyctereutes procyonoides (Gray, 1834) (see Yamamoto, 1994). Thus, Me. anakuma may be viewed as a facultative specialist of earthworms (a vermivore), although overall it is a generalist (Table 1). The diet diversity of the European badger Me. meles was shown to be negatively correlated with earthworm volume in the diet, assisting a facultative earthworm specialist property of “badgers” (Virgos et al., 2004; but see Roper, 1994 for the opportunistic generalist badger hypothesis).

      >  Mustela itatsi

    Because the Japanese weasel prefers wild rodents as a prey, it plays an important role as a biological pest control species (e.g., Uchida, 1969). Many individuals were released into many islands to suppress the rat density for an agricultural purpose (Masuda and Watanabe, 2009), leading to harmful influence on indigenous biota (Sekiguchi et al., 2002; Hamao et al., 2009). In contrast, previous feeding ecological studies have suggested that Mu. itatsi preys on a variety of vertebrates and invertebrates, and consumes fruits, depending on different seasons or regions (Yukawa, 1968a, 1968b; Uchida, 1969; Ohtsu, 1971; Asahi, 1975; Furuya et al., 1979; Fujii et al., 1998; Sekiguchi et al., 2002; Kaneko et al., 2009). Thus, this species is usually regarded as omnivorous (Masuda and Watanabe, 2009) and capable of adaptation to changes in food availability (Furuya et al., 1979; Fujii et al., 1998; Sekiguchi et al., 2002). However, the diet of this species seem to be mainly dominated by vertebrate animals if present as inferred from the aforementioned studies. Mu. itatsi is considered to be a facultative specialist of vertebrate animals (Table 1).

      >  Mustela sibirica

    The Siberian weasel is basically a flesh eater, mainly preying on rodents, shrews, and moles, but is able to adapt to a wide variety of food items including kitchen waste produced by human activities (Sasaki and Ono, 1994; Tatara and Doi, 1994; Nowak, 1999; Wu, 1999; Sasaki, 2009a). This species can be called an opportunistic predator of vertebrate animal and therefore a candidate competitor of the Japanese weasel Mu. itatsi (Table 1), although the former is more generalized to some extent. It is considered that the introduced Siberian weasel in the western Japan is expelling the indigenous Japanese weasel to more mountainous region probably because of the competitions on their ecological niche (Kuroda, 1955; Sasaki, 2009a).

      >  Mustela erminea

    The ermine (or the stoat) is an almost pure-carnivorous species, preying on small mammals (rodents and hares), birds, bird eggs, frogs, and insects (Nowak, 1999; Masuda, 2009a), and representing characteristics of a specialist predator (Table 1). Nevertheless, compared with Mu. nivalis, Mu. erminea has a more diverse diet (Elmeros, 2006). In an alpine habitat, it can even feed on fruits as a supplementary food in summer, probably to reduce the cost of capturing rodents as its preys (Martinoli et al., 2001). Mu. erminea is considered to have a more generalist property only in comparison with Mu. nivalis (King and Moors, 1979). Few studies about the diet of this species in Japan are available, except the studies of Ueuma and Tokuno (2001, 2002) and Ueuma et al. (2005) in Mt. Hakusan located in the nearly southern limit of the distribution in Honshu Islands. The authors showed that, as in continental populations of this species, the diet of Mu. erminea in Japan is mainly dominated by animal prey comprised of shrews, moles, rodents, hares, birds, and insects.

      >  Mustela nivalis

    The least weasel is also a nearly pure-carnivorous species (Nowak, 1999; Masuda, 2009b) (Table 1) and has a marked dietary overlap with Mu. erminea (Elmeros, 2006). Despite the overall similarity between Mu. erminea and Mu. nivalis in the prey category concentrated on small mammals and birds, the relative frequencies of their prey types differ; Mu. nivalis eats more small rodents while Mu. erminea adds more rabbits and birds as its main prey (King and Moors, 1979). In addition, these two species select different prey at the species level. For instance, in Denmark, Mu. erminea eats more Microtus voles and watervoles than Mu. nivalis, while Mu. nivalis preys on more bank voles and moles than Mu. erminea (Elmeros, 2006). The differences in food are probably mainly related to prey size, and this also concerns the differences in food of females and males within a species (King, 1989) Usually, Mu. nivalis is considered a more specialized carnivorous predator, exploiting efficiently small rodent preys than Mu. erminea because of its small body size (King and Moors, 1979). The diet of this species in Japan has not yet been thoroughly investigated.

    PHYLOGEOGRAPHIC AND FEEDING ECOLOGICAL EFFECTS ON MUSTELID FAUNAL ASSEMBLAGES IN JAPAN

    Community assembly and faunal organization could be affected by both phylogeographic and ecological causes. It is quite important to evaluate both ultimate (past historical) and proximate (present ecological) causes for the formation of the species assemblages (Losos, 1996; Webb et al., 2002; Wiens and Donoghue, 2004; Cardillo, 2011). In particular, the Japanese archipelagos possess an intrinsic complicated geo-history, such as different land connection patterns in different ages, and permit diverse ecological niche requirements in various landscapes, such as a long island structure extending from the northern subarctic to the southern subtropical climate zones. By elucidating processes of distributional patterns of mustelid species in Japan with extensive ecological diversity, it is possible to understand how the phylogeography and ecology concertedly contributes to the formation of the species assemblages.

    The aforementioned review of up-to-date studies on the origins and population history of the Japanese mustelid species and dietary niche breadth provide information that is useful in understanding the phylogeographic and feeding ecological reasons for the patterns of assemblages by mustelid carnivoran species in Japan. The answers to the following four questions revealed three different constraints that play important roles for the formation of the mustelid species distribution.

      >  Why is it possible for three mustelids endemic to Honshu-Shikoku-Kyushu Islands to be sympatric?

    Three species endemic to Honshu-Shikoku-Kyushu Islands, Martes melampus, Meles anakuma, and Mustela itatsi, which represent biogeographic pattern II, can be classified into different subfamilies (Fig. 3). Sato et al. (2012) adopted the framework of eight subfamilies within the family Mustelidae on the basis of their well-supported phylogeny, assigning Martes, Meles and Mustela to Guloninae, Melinae, and Mustelinae, respectively (see also Wolsan and Sato, 2010). The family Mustelidae is famous for their extensive ecomorphological diversifications at the subfamilial level (Fig. 3) (Sato et al., 2012; Wolsan, 2013). Species within Guloninae have mainly adapted to an arboreal life style and are adept at climbing trees. The Melinae species has large and strong claws for digging their den (“sett”), which has allowed the group members to adapt to a fossorial life style. The terrestrial Mustelinae species have adapted to more open and waterside habitats. Not surprisingly, dietary niche breadths of three species in different subfamilies are not largely overlapped (Table 1). Although not completely determined, Ma. melampus is a generalist, tasting diverse foods mainly from insects and vegetables, Me. anakuma is basically a specialist of earthworms, and Mu. itatsi mainly feed on vertebrate animals as a specialist (Table 1). Fundamental differences in their diet could have enabled them to co-occur on the Honshu-Shikoku- Kyushu Islands. It was previously implied that the coexistence of ecologically similar bird species was more difficult on islands than on mainland because of more extensive competition for limited resources in the islands (Grant, 1966). Niche heterogeneities could make such coexistence possible in the islands. It could be argued that the co-existence of three species in the Honshu-Shikoku-Kyushu Islands reflect pre-existing ancient ecological differences maintained by their evolutionary niche conservatism (Wiens and Graham, 2005). This is supportive evidence for the deep history hypothesis (present-day species assemblages was established largely due to ancient pre-existing ecological differences) rather than the competition hypothesis (present-day species assemblages was established largely due to ecological differences recently generated by competitions) as a primary cause for enabling the present-day species coexistence (Vitt and Pianka, 2005; also see Wu, 1999). The observation is also consistent with the trend that land-bridge island mammal assemblages tend to be composed of phylogenetically distantly related (overdispersed) lineages (Cardillo et al., 2008).

    Grant (1970) found a trend that islands are first occupied by generalists and then by specialists in the study of presence/ absence of some rodents on islands. Piechnik et al. (2008) also supported the generalist-before-specialist colonization hypothesis by analysing the colonization order of arthropods in islands. I tested the hypothesis by evaluating the phylogenetic history of three mustelid species endemic to Honshu- Shikoku-Kyushu Islands. The chronological inference showed that Ma. melampus, Me. anakuma, and Mu. itatsi originated 1.6-1.8, 0.5, and 1.5-1.6 Mya, respectively. Ages for intraspecific diversifications were estimated to be 0.32-0.36, 0.11-0.23, and 0.54-1.17 Mya, respectively. Although it is not clear when the migrations occurred after the time for the generation of the species, these estimates seemingly suggest the earliest colonization by Mu. itatsi, followed by Ma. melampus and Me. anakuma. The generalist-before-specialist hypothesis was therefore not clearly supported. This would probably be because not large overlap in their dietary niche permits any directions of the colonization order. The generalist- before-specialist rule may be applicable to organisms whose primary dietary niche is overlapped or nested, where the colonization of one of them are constrained by the other because of competitions for resources. The Taiwanese mustelids also shows similar tendency that lineages showing basic ecological differences colonized at different ages, independently of the generalist-before-specialist rule, in which the Ferret badgers Melogale moschata (Gray, 1831) in the subfamily Helictidinae migrated from the continent firstly, two Mustela species in Mustelinae secondly, and the Yellow- throated martens Martes flavigula (Boddaert, 1785) in Guloninae in the last (Hosoda et al., 2011). On the other hand, of two congeneric Mustela species in Taiwan whose primary diets such as rodents seem overlapped or nested, Mu. sibirica, a generalist, was inferred to be established earlier than Mu. nivalis, a specialist, which is consistent with the generalistbefore- specialist hypothesis. Also in Japan, two pure-carnivorous specialists, Mu. erminea and Mu. nivalis, seem to have been established later than the congeneric Mu. itatsi which is more generalized than Mu. erminea and Mu. nivalis. Collectively, an important suggestion could be drawn from the first question that evolutionary or phylogenetic constraints (pre-existing ecological differences) would strongly affect the species assemblage.

      >  Why are two Martes species allopatric in the Japanese archipelagos?

    Martes zibellina is present only in Hokkaido, not in Honshu, Shikoku, Kyushu and Tsushima Islands, whilst the natural distribution of Martes melampus is the reverse to that of Ma. zibellina (Fig. 2A, B), showing an allopatric distribution across the Tsugaru Strait located between Hokkaido and Honshu Islands (also known as Blakiston’s biogeographical line) (Fig. 1). Why is Ma. zibellina absent from southern islands despite of the possibility that Ma. zibellina representing more specialist property could colonize regions where Ma. melampus, a generalist, inhabits (the generalist-beforespecialist rule)? Why is Ma. melampus missing in Hokkaido Islands regardless of the possible ancestral migration from the northern route via Sakhalin and Hokkaido? Addressing these questions could provide us understandings of reasons for the allopatric distribution of two Martes species.

    According to the interspecific phylogenetic tree (Fig. 3) (Koepfli et al., 2008; Sato et al., 2012), the generation of the Ma. melampus lineage (1.6-1.8 Mya) occurred earlier than that of Ma. zibellina (ca. 1.0 Mya). It is natural to consider that the lineage of Ma. melampus first migrated into the Japanese archipelagos from the continent and then Ma. zibellina did. Based on the geological survey (Ohshima, 1990), the Tatar (Mamiya) and La Perouse (Soya) Straits (Fig. 1) could have been used as land bridges between the continent and Hokkaido through Sakhalin (Karafuto) in both the Middle (0.78-0.13 Mya) and Late (0.13-0.01 Mya) Pleistocene periods, while the Tsugaru Strait was only available for dispersals in the Middle Pleistocene but not in the Late Pleistocene. This is because decrease in sea level in Late Pleistocene was insufficient for land bridges to be formed in the Tsugaru Strait (Ohshima, 1990). Accordingly, it is possible to predict that Ma. melampus could have established their lineage in Honshu Islands before the onset of the Late Pleistocene, while Ma. zibellina could have migrated into Hokkaido in the Late Pleistocene. The time to the most recent common ancestor of examined individuals of Ma. melampus was estimated to be 0.32-0.36 Mya, which is in agreement with the Middle Pleistocene origin. Although time estimates for the origin of Ma. zibellina in Hokkaido were different in Sato et al. (2011), McKay (2012), and Ishida et al. (2013) adopting different data and chronological methods, they consistently showed that the migration could have occurred in the Late Pleistocene. Those estimates suggest that the arrival of the Ma. zibellina lineage in Hokkaido would have been too late to cross the Tsugaru Strait. This example of the allopatric distribution by two Martes species in Japan implies that “geological constraints” contributed largely to the formation of the species distribution.

    The answer for the second question is still not clear. However, Ishida et al. (2013) analyzed two closely linked nuclear genes of Ma. zibellina in Hokkaido and detected traces of ancient hybridization between Ma. zibellina and an ancestral lineage leading to Ma. melampus and Ma. americana, hinting at assimilation of the ancestral genomic components of Ma. melampus in the genome of Ma. zibellina. It is important to perform genome scale analyses for Ma. zibellina in Hokkaido to explore the reasons for disappearance of Ma. melampus linages from Hokkaido. This idea is conditional on the hypothesis that the migration of Ma. melampus via the nouthern route is true. However, the southern route hypothesis would not still be ruled out for the absence of Ma. melampus in Hokkaido.

      >  Why is the indigenous Siberian weasel absent in the Japanese archipelagos except Tsushima Islands?

    In spite of the inhabitation by Mustela sibirica in Tsushima Islands which is very closely located to the Japanese main islands, this species is not present in the main islands except the artificially introduced populations (Fig. 2D). Masuda et al. (2012) found specific lineages of Mu. sibirica in the Korean Peninsula, Taiwan, and the Tsushima Islands, and thus proposed the hypothesis that these regions would have played a role as refugia in the Pleistocene glacial periods to generate private mitochondrial DNA lineages. Considering that the current Korean and Tsushima Straits were formed at the similar age (Ohshima, 1990) and assuming that these straits would have probably been used as land bridges in the same glacial period of the Pleistocene, it is curious that Mu. sibirica is absent in the Japanese main islands despite of the presence in Tsushima Islands. Furthermore, the fact that Mu. sibirica is present in the Taiwan Islands increases the mystery. One explanation is that, as Masuda et al. (2012) supposed, the existence of Mu. itatsi might not have accepted the invasion of Mu. sibirica from the continent by competitive exclusion (Hardin, 1960; Cardillo et al., 2008). Although there are some differences in dietary habits and morphological characteristics pertaining to foraging strategy between these species (Yukawa, 1968a; Asahi, 1975; Suzuki et al., 2011), the dietary niche breadth are largely overlapped (Table 1) (Asahi, 1975). Therefore “ecological constraints (e.g., overlaps of their dietary niche)” would not allow two congeneric Mustela species to coexist in the Japanese main islands. It is not obvious whether the effect of inconsistency with the generalist-beforespecialist rule did not allow the migration of Mu. sibirica later than Mu. itatsi, where the former is slightly generalized than the latter. The other explanation for the absence of Mu. sibirica in the Japanese main islands may be associated with a geological constraint. Hosoda et al. (2000) suggested that the Cytb gene haplotype of Mu. sibirica in Tsushima Islands is the same as that observed in Korea. This implies that Korean Strait could have been used as a land bridge recently, while Tsushima Strait was not available. Concordantly, the leopard cat P. bengalensis in Tsushima Islands was shown to be very closely related to the continental Far East lineage, in which the divergence time was estimated to be ca. 0.03 Mya, while this species is not present in the Japanese main islands (Tamada et al., 2008). Furthermore, Mu. sibirica in Jeju Islands is also demonstrated to be not genetically differentiated from that in Korean mainland (Koh et al., 2012). Therefore eastern marginal islands around the Korean Peninsula could have been inhabited by Mu. sibirica in the geological situation that Tsushima Islands were not connected to the Japanese main islands. Thus, the geological constraint might have efficiently worked in the formation of the allopatric distributions of Mu. sibirica and Mu. itatsi, although further improvement of geological knowledge would be needed for Korean and Tsushima Straits.

      >  How do two pure-carnivorous Mustela species coexist in Japan?

    From the fossil evidence in Honshu Islands, Mu. erminea was found from layers of the Middle and Late Pleistocene periods, while Mu. nivalis was only from Late Pleistocene (Kawamura et al., 1989). Thus, it seems plausible that Mu. erminea was the earlier resident in Honshu than Mu. nivalis. The fact that the distribution of Mu. erminea is more expanded to the southern region than Mu. nivalis (Fig. 2E, F) (Masuda, 2009a, 2009b) may be the consequences of the establishment order. This idea agrees with the generalist-before-specialist hypothesis, because Mu. nivalis is considered more specialized to flesh preys than Mu. erminea (Elmeros, 2006). However, both genetic and morphological data have suggested that Mu. nivalis possesses more geographic variations than Mu. erminea across the Holarctic region (Abramov and Baryshnikov, 2000; Kurose et al., 2005), implying earlier diversifications among Mu. nivalis lineages than Mu. erminea. Therefore, in a future study, it will be necessary to assess more precise demographic history and the chronology of both species in order to evaluate which species migrated into the Japanese archipelagos first. Irrespective of the histories of both species, strict niche partitions might be needed for coexistence of the congeneric species whose niches are similar to each other. Available ecological data of these species in Japan are not sufficient to satisfactorily understand the mechanisms of their coexistence. However, if the ecological features are not extremely varied across the Eurasian continent and Japan, they can inhabit the same places by slightly different niche requirements. Dietary contents differ a bit from each other, as described above, where the niche breadth of Mu. erminea is slightly larger than Mu. nivalis (McDonald et al., 2000; Elmeros, 2006). Since Mu. nivalis is socially subordinate to dominant Mu. erminea (Erlinge and Sandell, 1988), Mu. nivalis would have to utilize resources that do not overlap with Mu. erminea (Morse, 1974). Mu. nivalis can more effectively catch small rodents by using its smaller body size and can breed more rapidly without delayed implantation (King and Moors, 1979). These differences in feeding and breeding strategies would reduce competitions between two species. In other words, “ecological constraints” may force two closely related species to divide their niche for their coexistence. Clearly further studies on phylogeography and dietary habits are needed to elucidate the historical and ecological causes for the assemblage of Mu. erminea and Mu. nivalis in Japan.

    CONCLUSION AND FUTURE PROBLEMS TO BE SOLVED

    Phylogeographic and ecological factors are both important facets in the species assemblage. By answering the four questions posed above, it was inferred that three constraints from evolutionary (or phylogenetic), geological, and ecological perspective have played effective roles in mustelid faunal assemblages in Japan. The large evolutionary difference tends to permit the coexistence without any major problems. Ecological constraints would force two closely related species to be allopatric by competitive exclusion or to be sympatric by resource partitions. Although these evolutionary and ecological constraints would have strong effects on the species assemblages, I suppose that the geological constraints would probably have the priority to generate the co-occurrence of species. If the geological situations do not permit species to be pooled in a particular place, it is obvious that the evolutionary or ecological causes would not work anymore. The mammalian faunal assemblage in the Japanese archipelagos would have been largely constrained by its complicated geohistory such as vicariance in the major straits in different ages. Geological effects would not be avoided in deciding which species is encompassed in the Japanese archipelagos. Further multidisciplinary studies would be needed to untangle the complicated effects on the faunal assemblage process leading to current species distributions.

    The phylogeographic and feeding ecological data are still insufficient to fully understand the relative strength of the aforementioned three factors. For further refinements of our knowledge about the importance of evolution (or phylogeny), geology, or ecology in the formation of the mustelid fauna in Japan, at least four issues need to be addressed. First, phylogeographic and demographic history of the resident species in Japan should be clarified with more fine scale time axis. Second, various niche spaces such as habitat preferences should be assessed in addition to dietary habits to examine niche breadth and overlap in more broad ecological sense. Interspecific interactions between introduced and native species in Japan (e.g., Mu. sibirica and Mu. itatsi in the western part of Honshu Islands) may provide valuable insight into the importance of the ecological constraint. Third, niche comparisons to carnivoran species other than mustelids occupying similar trophic positions would also be needed. For example, Me. anakuma in Mt. Nyugasa has a diet that is similar to Nyctereutes procyonoides (the raccoon dog) classified into different family (Canidae) than to Ma. melampus in the same family, possibly causing competitive interactions (Yamamoto, 1994; Koike et al., 2012). Therefore, considering only mustelids is insufficient to discuss the mechanisms of coexistence in the light of the ecological niche overlap and partition. Finally, the phylogenetic history of prey species should also be clarified to discuss the detail of the ecological constraints on the distribution of the Japanese mustelids. Clarifying phylogeography and ecology of an organism would satisfy the demand from the conservation biology that endemism and adaptation should both be considered (Ryder, 1986; de Guia and Saitoh, 2007; Isaac et al., 2007; Allendorf et al., 2010; Collen et al., 2011; Funk et al., 2012). Addressing the issues raised above will lead to appropriate conservation for small carnivoran species in one of the cradles of biodiversity in East Asia.

참고문헌
  • 1. Abramov AV 2002 Variation of the baculum structure of the Palaearctic badger (Carnivora, Mustelidae, Meles). [Russian Journal of Theriology] Vol.1 P.57-60 google
  • 2. Abramov AV 2003 The head colour pattern of the Eurasian badgers (Mustelidae, Meles). [Small Carnivore Conservation] Vol.29 P.5-7 google
  • 3. Abramov AV, Baryshnikov GF 2000 Geographic variation and intraspecific taxonomy of weasel Mustela nivalis (Carnivora, Mustelidae). [Zoosystematica Rossica] Vol.8 P.365-402 google
  • 4. Allendorf FW, Hohenlohe PA, Luikart G 2010 Genomics and the future of conservation genetics. [Nature Review Genetics] Vol.11 P.697-709 google cross ref
  • 5. Anderson E 1970 Quaternary evolution of the genus Martes (Carnivora, Mustelidae). [Acta Zoologica Fennica] Vol.130 P.1-132 google
  • 6. Arai S, Adachi T, Kuwahara Y, Yoshida K 2003 Food habit of the Japanese marten (Martes melampus) at Kuju highland in Kyushu, Japan. [Mammalian Science] Vol.43 P.19-28 google
  • 7. Asahi M 1975 Contents of alimentary canals of weasels collected in Kinki district, with a consideration of their caloric nutrition. [Zoological Magazine] Vol.84 P.190-195 google
  • 8. Cardillo M 2011 Phylogenetic structure of mammal assemblages at large geographical scales: linking phylogenetic community ecology with macroecology. [Philosophical Transactions of the Royal Society B Biological Sciences] Vol.366 P.2545-2553 google cross ref
  • 9. Cardillo M, Gittleman JL, Purvis A 2008 Global patterns in the phylogenetic structure of island mammal assemblages. [Proceedings of the Royal Society of London Series B Biological Sciences] Vol.275 P.1549-1556 google cross ref
  • 10. Cerro ID, Marmi J, Ferrando A, Chashchin P, Taberlet P, Bosch M 2010 Nuclear and mitochondrial phylogenies provide evidence for four species of Eurasian badgers (Carnivora). [Zoologica Scripta] Vol.39 P.415-425 google cross ref
  • 11. Collen B, Turvey ST, Waterman C, Meredith HMR, Kuhn TS, Baillie JEM, Isaac NJB 2011 Investing in evolutionary history: implementing a phylogenetic approach for mammal conservation. [Philosophical Transactions of the Royal Society B Biological Sciences] Vol.366 P.2611-2622 google cross ref
  • 12. de Guia APO, Saitoh T 2007 The gap between the concept and definitions in the evolutionarily significant unit: the need to integrate neutral genetic variation and adaptive variation. [Ecological Research] Vol.22 P.604-612 google cross ref
  • 13. Dobson M 1994 Patterns of distribution in Japanese land mammals. [Mammal Review] Vol.24 P.91-111 google cross ref
  • 14. Dobson M, Kawamura Y 1998 Origin of the Japanese land mammal fauna: allocation of extant species to historicallybased categories. [The Quaternary Research] Vol.37 P.385-395 google cross ref
  • 15. Elmeros M 2006 Food habits of stoats Mustela erminea and weasels Mustela nivalis in Denmark. [Acta Theriologica] Vol.51 P.179-186 google cross ref
  • 16. Erlinge S, Sandell M 1988 Coexistence of stoat, Mustela erminea, and weasel, M. nivalis: social dominance, scent communication, and reciprocal distribution. [Oikos] Vol.53 P.242-246 google cross ref
  • 17. Frankham R, Ballou JD, Briscoe DA 2010 Introduction to conservation genetics. P.1-642 google
  • 18. Fu YX 1997 Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. [Genetics] Vol.147 P.915-925 google
  • 19. Fujii T, Maruyama N, Kanzaki N 1998 Seasonal changes in food habits of Japanese weasel in a middle stream of the Tamagawa River. [Mammalian Science] Vol.38 P.1-8 google
  • 20. Funk WC, McKay JK, Hohenlohe PA, Allendorf FW 2012 Harnessing genomics for delineating conservation units. [Trends in Ecology and Evolution] Vol.27 P.489-496 google cross ref
  • 21. Furuya Y, Kishida R, Sendo K, Noguchi K, Yamasaki M 1979 Seasonal changes of food habit of weasels (Mustela sibirica) in Nishikuma Valley, Kochi Prefecture. [The Journal of the Mammalogical Society of Japan] Vol.8 P.1-11 google
  • 22. Grant PR 1966 Ecological compatibility of bird species on islands. [The American Naturalist] Vol.100 P.451-462 google cross ref
  • 23. Grant PR 1970 Colonization of islands by ecologically dissimilar species of mammals. [Canadian Journal of Zoology] Vol.48 P.545-553 google cross ref
  • 24. Hamao S, Nishimatsu K, Kamito T 2009 Predation of bird nests by introduced Japanese Weasel Mustela itatsi on an island. [Ornithological Science] Vol.8 P.139-146 google cross ref
  • 25. Hardin G 1960 The competitive exclusion principle. [Science] Vol.131 P.1292-1297 google cross ref
  • 26. Hosoda T, Sato JJ, Lin LK, Chen YJ, Harada M, Suzuki H 2011 Phylogenetic history of mustelid fauna in Taiwan inferred from mitochondrial genetic loci. [Canadian Journal of Zoology] Vol.89 P.559-569 google cross ref
  • 27. Hosoda T, Suzuki H, Harada M, Tsuchiya K, Han SH, Zhang YP, Kryukov AP, Lin LK 2000 Evolutionary trends of the mitochondrial lineage differentiation in species of genera Martes and Mustela. [Genes & Genetic Systems] Vol.75 P.259-267 google
  • 28. Hosoda T, Suzuki H, Iwasa MA, Hayashida M, Watanabe S, Tatara M, Tsushiya K 1999 Genetic relationships within and between the Japanese marten Martes melampus and the sable M. zibellina, based on variation of mitochondrial DNA and nuclear ribosomal DNA. [Mammal Study] Vol.24 P.25-33 google cross ref
  • 29. Inoue T, Murakami T, Abramov AV, Masuda R 2010 Mitochondrial DNA control region variations in the sable Martes zibellina of Hokkaido Island and the Eurasian continent, compared with the Japanese marten M. melampus. [Mammal Study] Vol.35 P.145-155 google cross ref
  • 30. Isaac NJB, Turvey ST, Collen B, Waterman C, Baillie JEM 2007 Mammals on the EDGE: conservation priorities based on threat and phylogeny. [PLoS ONE] Vol.2 P.e296 google cross ref
  • 31. Ishida K, Sato JJ, Kinoshita G, Hosoda T, Kryukov AP, Suzuki H 2013 Evolutionary history of the sable (Martes zibellina brachyura) on Hokkaido inferred from mitochondrial Cytb and nuclear Mc1r and Tcf25 gene sequences. [Acta Theriologica] Vol.58 P.13-24 google cross ref
  • 32. Kaneko Y 2009 Meles anakuma Temminck, 1844. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.258-260 google
  • 33. Kaneko Y, Maruyama N, Macdonald DW 2006 Food habits and habitat selection of suburban badgers (Meles meles) in Japan. [Journal of Zoology] Vol.270 P.78-89 google
  • 34. Kaneko Y, Shibuya M, Yamaguchi N, Fujii T, Okumura T, Matsubayashi K, Hioki Y 2009 Diet of Japanese weasels (Mustela itatsi) in a sub-urban landscape: implications for year-round persistence of local populations. [Mammal Study] Vol.34 P.97-105 google cross ref
  • 35. Kawamura Y 1991 Quaternary mammalian faunas in the Japanese Islands. [The Quaternary Research] Vol.30 P.213-220 google cross ref
  • 36. Kawamura Y 1994 Late Pleistocene to Holocene mammalian faunal succession in the Japanese Islands, with comments on the Late Quaternary extinctions. [Archaeozoologia] Vol.6 P.7-22 google
  • 37. Kawamura Y, Kamei T, Taruno H 1989 Middle and Late Pleistocene mammalian faunas in Japan. [The Quaternary Research] Vol.28 P.317-326 google cross ref
  • 38. Kier G, Kreft H, Lee TM, Jetz W, Ibisch PL, Nowicki C, Mutke J, Barthlott W 2009 A global assessment of endemism and species richness across island and mainland regions. [Proceedings of the National Academy of Sciences of the United States of America] Vol.106 P.9322-9327 google cross ref
  • 39. King CM 1989 The advantages and disadvantages of small size to weasels, Mustela species. In: Carnivore behavior, ecology, and evolution. Vol. 1 (Ed., Gittleman JL). P.302-334 google
  • 40. King CM, Moors PJ 1979 On co-existence, foraging strategy and the biogeography of weasels and stoats (Mustela nivalis and M. erminea) in Britain. [Oecologia] Vol.39 P.129-150 google cross ref
  • 41. Koepfli KP, Deere KA, Slater GJ, Begg C, Begg K, Grassman L, Lucherini M, Veron G, Wayne RK 2008 Multigene phylogeny of the Mustelidae: resolving relationships, tempo and biogeographic history of a mammalian adaptive radiation. [BMC Biology] Vol.6 P.10 google cross ref
  • 42. Koh HS, Jang KH, Oh JG, Han ED, Jo JE, Ham EJ, Jeong SK, Lee JH, Kim KS, Kweon GH, In ST 2012 Lack of mitochondrial DNA sequence divergence between two subspecies of the Siberian weasel from Korea: Mustela sibirica coreanus from the Korean Peninsula and M. s. quelpartis from Jeju Island. [Animal Systematics, Evolution and Diversity] Vol.28 P.133-136 google cross ref
  • 43. Koike S, Morimoto H, Goto Y, Kozakai C, Yamazaki K 2012 Insectivory by five sympatric carnivores in cool-temperate deciduous forests. [Mammal Study] Vol.37 P.73-83 google cross ref
  • 44. Kuroda N 1955 The present status of the introduced mammals in Japan. [The Journal of the Mammalogical Society of Japan] Vol.1 P.13-18 google
  • 45. Kuroda N, Mori T 1923 Two new and rare mammals from Korea. [Journal of Mammalogy] Vol.4 P.27-28 google cross ref
  • 46. Kurose N, Abramov AV, Masuda R 2005 Comparative phylogeography between the ermine Mustela erminea and the least weasel M. nivalis of Palaearctic and Nearctic regions, based on analysis of mitochondrial DNA control region sequences. [Zoological Science] Vol.22 P.1069-1078 google cross ref
  • 47. Kurose N, Kaneko Y, Abramov AV, Siriaroonrat B, Masuda R 2001 Low genetic diversity in Japanese populations of the Eurasian badger Meles meles (Mustelidae, Carnivora) revealed by mitochondrial cytochrome b gene sequences. [Zoological Science] Vol.18 P.1145-1151 google cross ref
  • 48. Kurose N, Masuda R, Siriaroonrat B, Yoshida MC 1999 Intraspecific variation of mitochondrial cytochrome b gene sequences of the Japanese marten Martes melampus and the Sable Martes zibellina (Mustelidae, Carnivora, Mammalia) in Japan. [Zoological Science] Vol.16 P.693-700 google cross ref
  • 49. Losos JB 1996 Phylogenetic perspectives on community ecology. [Ecology] Vol.77 P.1344-1354 google cross ref
  • 50. Marmi J, Lopez-Giraldez JF, Domingo-Roura X 2004 Phylogeny, evolutionary history and taxonomy of the Mustelidae based on sequences of the cytochrome b gene and a complex repetitive flanking region. [Zoologica Scripta] Vol.33 P.481-499 google cross ref
  • 51. Marmi J, Lopez-Giraldez F, Macdonald DW, Calafell F, Zholnerovskaya E, Domingo-Roura X 2006 Mitochondrial DNA reveals a strong phylogeographic structure in the badger across Eurasia. [Molecular Ecology] Vol.15 P.1007-1020 google cross ref
  • 52. Martinoli A, Preatoni DG, Chiarenzi B, Wauters LA, Tosi G 2001 Diet of stoats (Mustela erminea) in an Alpine habitat: the importance of fruit consumption in summer. [Acta Oecologica] Vol.22 P.45-53 google cross ref
  • 53. Masuda R 2009a Mustela erminea Linnaeus, 1758. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.246-247 google
  • 54. Masuda R 2009b Mustela nivalis Linnaeus, 1766. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.244-245 google
  • 55. Masuda R, Kurose N, Watanabe S, Abramov AV, Han SH, Lin LK, Oshida T 2012 Molecular phylogeography of the Japanese weasel, Mustela itatsi (Carnivora: Mustelidae), endemic to the Japanese islands, revealed by mitochondrial DNA analysis. [Biological Journal of the Linnean Society] Vol.107 P.307-321 google cross ref
  • 56. Masuda R, Watanabe S 2009 Mustela itatsi Temminck, 1844. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.240-241 google
  • 57. Masuda R, Yoshida MC 1994 A molecular phylogeny of the family Mustelidae (Mammalia, Carnivora), based on comparison of mitochondrial cytochrome b nucleotide sequences. [Zoological Science] Vol.11 P.605-612 google
  • 58. McDonald RA, Webbon C, Harris S 2000 The diet of stoats (Mustela erminea) and weasels (Mustela nivalis) in Great Britain. [Journal of Zoology] Vol.252 P.363-371 google cross ref
  • 59. McKay BD 2012 A new timeframe for the diversification of Japan’s mammals. [Journal of Biogeography] Vol.39 P.1134-1143 google cross ref
  • 60. Millien V 2006 Morphological evolution is accelerated among island mammals. [PLoS Biology] Vol.4 P.e321 google cross ref
  • 61. Millien-Parra V, Jaeger JJ 1999 Island biogeography of the Japanese terrestrial mammal assemblages: an example of a relict fauna. [Journal of Biogeography] Vol.26 P.959-972 google cross ref
  • 62. Monakhov VG 2011 Martes zibellina (Carnivora: Mustelidae). [Mammalian Species] Vol.43 P.75-86 google cross ref
  • 63. Morse DH 1974 Niche breadth as a function of social dominance. [The American Naturalist] Vol.108 P.818-830 google cross ref
  • 64. Murakami T 2003 Food habits of the Japanese sable Martes zibellina brachyura in eastern Hokkaido, Japan. [Mammal Study] Vol.28 P.129-134 google cross ref
  • 65. Murakami T 2009 Martes zibellina (Linnaeus, 1758). In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.252-253 google
  • 66. Nakamura T, Kanzaki N, Maruyama N 2001 Seasonal changes in food habits of Japanese martens in Hinode-cho and Akiruno- shi, Tokyo. [Wildlife Conservation Japan.] Vol.6 P.15-24 google
  • 67. Nowak RM 1999 Walker’s mammals of the world. Vol. 1. P.1-1936 google
  • 68. Obara Y 1991 Karyosystematics of the mustelid carnivores of Japan. [Mammalian Science] Vol.30 P.197-220 google
  • 69. Ogino S, Otsuka H, Harunari H 2009 The middle Pleistocene Matsugae fauna, northern Kyushu, West Japan. [Paleontological Research] Vol.13 P.367-384 google cross ref
  • 70. Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T 2009 The wild mammals of Japan. P.1-544 google
  • 71. Ohshima K 1990 The history of straits around the Japanese Islands in the Late-Quaternary. [The Quaternary Research] Vol.29 P.193-208 google cross ref
  • 72. Ohtsu M 1971 Food habits of the Japanese weasel in winter and its conservation. [Journal of Japanese Society of Applied Entomology and Zoology] Vol.15 P.87-88 google cross ref
  • 73. Otani T 2002 Seed dispersal by Japanese marten Martes melampus in the subalpine shrubland of northern Japan. [Ecological Research] Vol.17 P.29-38 google cross ref
  • 74. Piechnik DA, Lawler SP, Martinez ND 2008 Food-web assembly during a classic biogeographic study: species’ “trophic breadth” corresponds to colonization order. [Oikos] Vol.117 P.665-674 google cross ref
  • 75. Ricketts TH, Dinerstein E, Boucher T, Brooks TM, Butchart SHM, Hoffmann M, Lamoreux JF, Morrison J, Parr M, Pilgrim JD, Rodrigues ASL, Sechrest W, Wallace GE, Berlin K, Bielby J, Bielby ND, Church DR, Cox N, Knox D, Loucks C, Luck GW, Master LL, Moore R, Naidoo R, Ridgely R, Schatz GE, Shire G, Strand H, Wettengel W, Wikramanayake E 2005 Pinpointing and preventing imminent extinctions. [Proceedings of the National Academy of Sciences of the United States of America] Vol.102 P.18497-18501 google cross ref
  • 76. Roper TJ 1994 The European badger Meles meles: food specialist or generalist? [Journal of Zoology, London] Vol.234 P.437-452 google cross ref
  • 77. Ryder OA 1986 Species conservation and systematics: the dilemma of subspecies. [Trends in Ecology and Evolution] Vol.1 P.9-10 google cross ref
  • 78. Sasaki H 2009a Mustela sibirica Pallas, 1773. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.242-243 google
  • 79. Sasaki H 2009b Lutra lutra (Linnaeus, 1758). In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.254-255 google
  • 80. Sasaki H, Ono Y 1994 Habitat use and selection of the Siberian weasel Mustela sibirica coreana during the non-mating season. [Journal of the Mammalogical Society of Japan] Vol.19 P.21-32 google
  • 81. Sato JJ, Hosoda T, Kryukov AP, Kartavtseva IV, Suzuki H 2011 Genetic diversity of the sable (Martes zibellina, Mustelidae) in Russian Far East and Hokkaido inferred from mitochondrial NADH dehydrogenase subunit 2 gene sequences. [Mammal Study] Vol.36 P.209-222 google cross ref
  • 82. Sato JJ, Hosoda T, Wolsan M, Suzuki H 2004 Molecular phylogeny of arctoids (Mammalia; Carnivora) with emphasis on phylogenetic and taxonomic positions of the ferret-badgers and skunks. [Zoological Science] Vol.21 P.111-118 google cross ref
  • 83. Sato JJ, Hosoda T, Wolsan M, Tsuchiya K, Yamamoto Y, Suzuki H 2003 Phylogenetic relationships and divergence time among mustelids (Mammalia; Carnivora) based on nucleotide sequences of the nuclear interphotoreceptor retinoid binding protein and mitochondrial cytochrome b genes. [Zoological Science] Vol.20 P.243-264 google cross ref
  • 84. Sato JJ, Wolsan M, Minami S, Hosoda T, Sinaga MH, Hiyama K, Yamaguchi Y, Suzuki H 2009a Deciphering and dating the red panda’s ancestry and early adaptive radiation of Musteloidea. [Molecular Phylogenetics and Evolution] Vol.53 P.907-922 google cross ref
  • 85. Sato JJ, Wolsan M, Prevosti FJ, D’Eelia G, Begg C, Begg K, Hosoda T, Campbell KL, Suzuki H 2012 Evolutionary and biogeographic history of weasel-like carnivorans (Musteloidea). [Molecular Phylogenetics and Evolution] Vol.63 P.745-757 google cross ref
  • 86. Sato JJ, Wolsan M, Suzuki H, Hosoda T, Yamaguchi Y, Hiyama K, Kobayashi M, Minami S 2006 Evidence from nuclear DNA sequences shed light on the phylogenetic relationships of Pinnipedia: single origin with affinity to Musteloidea. [Zoological Science] Vol.23 P.125-146 google cross ref
  • 87. Sato JJ, Yasuda SP, Hosoda T 2009b Genetic diversity of the Japanese marten (Martes melampus) and its implications for the conservation unit. [Zoological Science] Vol.26 P.457-466 google cross ref
  • 88. Sekiguchi K, Ogura G, Sasaki T, Nagayama Y, Tsuha K, Kawashima Y 2002 Food habits of introduced Japanese weasels (Mustela itatsi) and impacts on native species on Zamami Island. [Mammalian Science] Vol.42 P.153-160 google
  • 89. Shiratsuki N 1972 On the feces of the Japanese marten. [Nature Study] Vol.18 P.98-101 google
  • 90. Suzuki H 2009 A molecular phylogenetic view of mammals in the “three-story museum” of Hokkaido, Honshu, and Ryukyu Islands, Japan. In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.261-263 google
  • 91. Suzuki S, Abe M, Motokawa M 2011 Allometric comparison of skulls from two closely related weasels, Mustela itatsi and M. sibirica. [Zoological Science] Vol.28 P.676-688 google cross ref
  • 92. Suzuki S, Miyao T, Nishizawa T, Shida Y, Takada Y 1976 Studies on mammals of the Mt. Koso-Komagatake, central Japan Alps. II. Food habit of the Japanese marten in autumn and winter in upper part of low mountainous zone on eastern slope of the Kiso-Komagatake, with special reference to the scat analysis. [Journal of the Faculty of Agriculture Shinshu University] Vol.13 P.21-42 google
  • 93. Suzuki S, Miyao T, Nishizawa T, Takada Y 1977 Studies on mammals of the Mt. Koso-Komagatake, central Japan Alps. III. Food habit of the Japanese marten in upper part of low mountainous zone and sub-alpine zone of the Kiso-Komagatake. [Journal of the Faculty of Agriculture Shinshu University] Vol.14 P.147-178 google
  • 94. Tajima F 1989 Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. [Genetics] Vol.123 P.585-595 google
  • 95. Tamada T, Siriaroonrat B, Subramaniam V, Hamachi M, Lin LK, Oshida T, Rerkamnuaychoke W, Masuda R 2008 Molecular diversity and phylogeography of the Asian leopard cat, Felis bengalensis, inferred from mitochondrial and Y-chromosomal DNA sequences. [Zoological Science] Vol.25 P.154-163 google cross ref
  • 96. Tashima S, Kaneko Y, Anezaki T, Baba M, Yachimori S, Abramov AV, Saveljev AP, Masuda R 2011 Phylogeographic sympatry and isolation of the Eurasian badgers (Meles, Mustelidae, Carnivora): implications for an alternative analysis using maternally as well as paternally inherited genes. [Zoological Science] Vol.28 P.293-303 google cross ref
  • 97. Tatara M, Doi T 1994 Comparative analyses on food habits of Japanese marten, Siberian weasel and leopard cat in the Tsushima islands, Japan. [Ecological Research] Vol.9 P.99-107 google cross ref
  • 98. Tsuji Y, Tatewaki T, Kanda E 2011 Endozoochorous seed dispersal by sympatric mustelids, Martes melampus and Mustela itatsi, in western Tokyo, central Japan. [Mammalian Biology] Vol.76 P.628-633 google
  • 99. Uchida T 1969 Rat-control procedures on the Pacific islands, with special reference to the efficiency of biological control agents. II: Efficiency of the Japanese weasel, Mustela sibirica itatsi Temminck & Schlegel, as a rat-control agent in the Ryukyus. [Journal of the Faculty of Agriculture, Kyushu University] Vol.15 P.355-385 google
  • 100. Ueuma Y, Tokuno C 2001 Droppings of red fox (Vulpes vulpes japonica), Japanese marten (Martes melampus melampus), and Hondo stoat (Mustela erminea nippon) on the alpine trails in Mt. Hakusan. [Research Reports of Hakusan Nature Conservation Center in Ishikawa Prefecture] Vol.28 P.7-11 google
  • 101. Ueuma Y, Tokuno C 2002 Droppings of red fox (Vulpes vulpes japonica), Japanese marten (Martes melampus melampus), and Hondo stoat (Mustela erminea nippon) on the alpine trails in Mt. Hakusan in 2002. [Research Reports of Hakusan Nature Conservation Center in Ishikawa Prefecture] Vol.29 P.55-58 google
  • 102. Ueuma Y, Tokuno C, Tsuji M 2005 Food of red fox (Vulpes vulpes japonica), Japanese marten (Martes melampus melampus), and Hondo stoat (Mustela erminea nippon) analysed by dropping contents on the trails in Mt. Hakusan. [Research Reports of Hakusan Nature Conservation Center in Ishikawa Prefecture] Vol.32 P.31-36 google
  • 103. Uraguchi K 2009 Neovison vison (Schreber, 1777). In: The wild mammals of Japan (Eds., Ohdachi SD, Ishibashi Y, Iwasa MA, Saitoh T). P.248-249 google
  • 104. Virgos E, Mangas J, Blanco-Aguiar JA, Garrote G, Almagro N, Viso RP 2004 Food habits of European badgers (Meles meles) along an altitudinal gradient of Mediterranean environments: a field test of the earthworm specialization hypothesis. [Canadian Journal of Zoology] Vol.82 P.41-51 google cross ref
  • 105. Vitt LJ, Pianka ER 2005 Deep history impacts present-day ecology and biodiversity. [Proceedings of the National Academy of Sciences of the United States of America] Vol.102 P.7877-7881 google cross ref
  • 106. Wiens JJ, Donoghue MJ 2004 Historical biogeography, ecology and species richness. [Trends in Ecology and Evolution] Vol.19 P.639-644 google cross ref
  • 107. Wiens JJ, Graham CH 2005 Niche conservatism: integrating evolution, ecology, and conservation biology. [Annual Review of Ecology, Evolution, and Systematics] Vol.36 P.519-539 google cross ref
  • 108. Webb CO, Ackerly DD, McPeek MA, Donoghue MJ 2002 Phylogenies and community ecology. [Annual Review of Ecology and Systematics] Vol.33 P.475-505 google cross ref
  • 109. Wolsan M 2013 Family Mustelidae: weasels, polecats, otters, Ratel and allies. In: Mammals of Africa. Vol. 5. Carnivores, pangolins, equids and rhinoceroses (Eds., Kingdon J, Hoffmann M). P.82-84 google
  • 110. Wolsan M, Sato JJ 2010 Effects of data incompleteness on the relative performance of parsimony and Bayesian approaches in a supermatrix phylogenetic reconstruction of Mustelidae and Procyonidae (Carnivora). [Cladistics] Vol.26 P.168-194 google cross ref
  • 111. Wozencraft WC 1993 Order Carnivora. In: Mammal species of the world: a taxonomic and geographic reference. (Eds., Wilson DE, Reeder DM). P.279-348 google
  • 112. Wozencraft WC 2005 Order Carnivora. In: Mammal species of the world: a taxonomic and geographic reference. (Eds., Wilson DE, Reeder DM). P.532-628 google
  • 113. Wu HY 1999 Is there current competition between sympatric Siberian weasels (Mustela sibirica) and ferret badgers (Melogale moschata) in a subtropical forest ecosystem of Taiwan? [Zoological Studies] Vol.38 P.443-451 google
  • 114. Yamagishi M 1990 Seasonal food habits of the Japanese marten. [Bulletin of the Tokyo University Forests] Vol.83 P.9-18 google
  • 115. Yamamoto Y 1991 Food habits of the Japanese badger (Meles meles anakuma) in Mt. Nyugasa, Nagano Prefecture, Japan. [The Natural Environmental Science Research] Vol.4 P.73-83 google
  • 116. Yamamoto Y 1994 Comparative analyses on food habits of Japanese marten, red fox, badger and raccoon dog in Mt. Nyugasa, Nagano Prefecture, Japan. [The Natural Environmental Science Research] Vol.7 P.45-52 google
  • 117. Yukawa M 1968a Notes on some ecology and measurements of Mustela sibirica coreana Domaniewski. [The Journal of the Mammalogical Society of Japan] Vol.4 P.58-59 google
  • 118. Yukawa M 1968b On the food habit of the Japanese weasel (Mustela sibirica itatsi Temminck) in northern parts of Hiroshima Pref. [Hiwa Science Museum Report] Vol.12 P.7-10 google
  • 119. Zhou YB, Newman C, Xu WT, Buesching CD, Zalewski A, Kaneko Y, Macdonald DW, Xie ZQ 2011 Biogeographical variation in the diet of Holarctic martens (genus Martes, Mammalia: Carnivora: Mustelidae): adaptive foraging in generalists. [Journal of Biogeography] Vol.38 P.137-147 google cross ref
OAK XML 통계
이미지 / 테이블
  • [ Fig. 1. ]  A diagram for configurations of islands around the Japanese archipelagos (Sakhalin [Karafuto], Hokkaido, Honshu, Shikoku, Kyushu, Tsushima, Ryukyu, Jeju, and Taiwan Islands) and major straits (Tatar [Mamiya], La Perouse [Soya], Tsugaru, Tsushima, Korean, Tokara, and Kerama Straits).
    A diagram for configurations of islands around the Japanese archipelagos (Sakhalin [Karafuto], Hokkaido, Honshu, Shikoku, Kyushu, Tsushima, Ryukyu, Jeju, and Taiwan Islands) and major straits (Tatar [Mamiya], La Perouse [Soya], Tsugaru, Tsushima, Korean, Tokara, and Kerama Straits).
  • [ Fig. 2. ]  Natural distributions of six extant and indigenous terrestrial mustelid species inhabiting the Japanese archipelagos. A, Martes zibellina (black); B, Martes melampus (black); C, Meles anakuma (black); D, Mustela itatsi (black) and Mustela sibirica (grey); E, Mustela erminea (black); F, Mustela nivalis (black). I referred Ohdachi et al. (2009) and Wozencraft (2005) for information of their distributions. However, the southern limit of the continental distributions of Ma. zibellina, Mu. erminea, Mu. nivalis and precise distribution of Mu. erminea and Mu. nivalis in Honshu Islands should be regarded as provisional information.
    Natural distributions of six extant and indigenous terrestrial mustelid species inhabiting the Japanese archipelagos. A, Martes zibellina (black); B, Martes melampus (black); C, Meles anakuma (black); D, Mustela itatsi (black) and Mustela sibirica (grey); E, Mustela erminea (black); F, Mustela nivalis (black). I referred Ohdachi et al. (2009) and Wozencraft (2005) for information of their distributions. However, the southern limit of the continental distributions of Ma. zibellina, Mu. erminea, Mu. nivalis and precise distribution of Mu. erminea and Mu. nivalis in Honshu Islands should be regarded as provisional information.
  • [ Fig. 3. ]  Interspecific phylogenetic relationships and divergence times of lineages within the family Mustelidae reconstructed on the basis of recent molecular phylogenetic studies (Sato et al., 2003, 2004, 2006, 2009a, 2012; Koepfli et al., 2008; Wolsan and Sato, 2010; Tashima et al., 2011). I followed Wolsan and Sato (2010) and Sato et al. (2012) for the subfamilial designation. Branches and scientific names for the Japanese mustelid species were highlighted with thick line and bold font, respectively. I assigned ecological types (arboreal, aquatic, fossorial, and terrestrial) to each subfamily according to Nowak (1999), although some exceptions exist (e.g., semi-aquatic Neovison vison in Mustelinae). Mya, million years ago.
    Interspecific phylogenetic relationships and divergence times of lineages within the family Mustelidae reconstructed on the basis of recent molecular phylogenetic studies (Sato et al., 2003, 2004, 2006, 2009a, 2012; Koepfli et al., 2008; Wolsan and Sato, 2010; Tashima et al., 2011). I followed Wolsan and Sato (2010) and Sato et al. (2012) for the subfamilial designation. Branches and scientific names for the Japanese mustelid species were highlighted with thick line and bold font, respectively. I assigned ecological types (arboreal, aquatic, fossorial, and terrestrial) to each subfamily according to Nowak (1999), although some exceptions exist (e.g., semi-aquatic Neovison vison in Mustelinae). Mya, million years ago.
  • [ Table 1. ]  Dietary habits of seven extant mustelid species in Japan
    Dietary habits of seven extant mustelid species in Japan
(우)06579 서울시 서초구 반포대로 201(반포동)
Tel. 02-537-6389 | Fax. 02-590-0571 | 문의 : oak2014@korea.kr
Copyright(c) National Library of Korea. All rights reserved.